Introduction

Owing to the fast development of electric vehicles and grid energy storage, there is increasing demand for lithium-ion batteries with long cycle life1,2,3. As compared to the negative electrode material of graphite, the positive electrode materials possess relatively lower specific capacity and shorter cycle life, which primarily limit the electrochemical performance of lithium-ion batteries4,5,6. To achieve high energy density and long cycle life at the same time, the positive electrode materials need to be able to accommodate a large amount of lithium ions without imposing permanent change to the crystal structure. Large lattice variations accompanied with Li+ intercalation/deintercalation will degrade the structure of positive electrode materials, which could induce irreversible phase transition and shorten cycle life. Especially for the solid-state lithium batteries, large volume variation in the positive electrode materials could lead to a breach of the electrode–solid electrolyte interface, resulting in large interfacial impedance and battery failure7,8. However, there is always a trade-off between large specific capacity and cycle life for positive electrode materials, and developing stabilized intercalation positive electrodes that can accommodate a large quantity of lithium ions is still a challenge.

Among the various positive electrode materials, layered α-MoO3 is particularly appealing to lithium ion insertion due to its edge-sharing MoO6 octahedra bilayers stacked along [010] by van der Waals (vdW) interaction9. A high theoretical specific capacity of about 372 mAh g−1 is expected, corresponding to the Mo6+/Mo4+ redox couple with two lithium storage per Mo, which makes α-MoO3 a promising positive electrode for Li-ion storage10,11,12. However, the insertion of Li+ ions into the layered α-MoO3 breaks the thermodynamically stable state and induces large lattice expansion (~16%), resulting in the irreversible phase transformation to acicular LixMoO3 (x ~0.25) and fast capacity fading (Fig. 1a)13,14. To prolong the cycle life of α-MoO3, various strategies including constructing nanostructures15, nitrogen doping16, and widening the vdW interlayer10,13 of α-MoO3 have been extensively explored. Nevertheless, there is still notable capacity loss, especially during the initial cycles for the modified α-MoO310,13, suggesting more effective approaches need to be developed to further stabilize the layered structure without compromising its large specific capacity.

Fig. 1: Coherent interface design to suppress lattice change of layered MoO3 during lithiation.
figure 1

Schematic illustrations of a large lattice expansion of the individual α-MoO3, b minimal lattice expansion of the individual η-Mo4O11, and c suppressed lattice expansion of α-MoO3 via coherent interface in the α-MoO3/η-Mo4O11 heterostructure during lithiation.

Recent studies on constructing spinel-layered heterostructures for LiMnO2 and Li-rich (Li2MnO3) electrode materials suggest that interface engineering could be an effective way to improve the structural stability of Li+ intercalation host17,18. To reduce the large lattice expansion of α-MoO3 during the lithiation process, we propose a coherent epitaxial interface pinning design, which can inhibit the mobility of Mo atoms via strong ionic bonds at coherent interfaces. In this study, we constructed α-MoO3/η-Mo4O11 intergrown heterostructure with coherent epitaxial interface by magnetron sputtering. η-Mo4O11 is a nearly “zero-strain” Li+ intercalation host (Fig. 1b), which could be the perfect structural stabilizer for the coherent interface design (Fig. 1c). When a coherent interface can be established between α-MoO3 and η-Mo4O11, the high structural stability of η-Mo4O11 can be transferred to α-MoO3, which could suppress lattice variation of α-MoO3 during lithiation and delithiation. Therefore, the coherent interface pinning design could provide an effective structural regulation strategy to stabilize electrode materials with intercalation chemistry.

Herein, pure α-MoO3, α-MoO3/η-Mo4O11, and pure η-Mo4O11 thin films were prepared respectively by manipulating the O2/Ar flow ratio during the sputtering process. With the coherent interface, the large lattice expansion and phase transition in α-MoO3 have been viably suppressed during the lithiation, which was evidenced by the ex situ XRD. The density functional theory (DFT) calculations indicate the interface pinning can effectively improve the structural stability of α-MoO3 and suppress Mo migration during charge/discharge. Consequently, the α-MoO3/η-Mo4O11 heterostructured electrode achieved a reversible specific capacity as high as ~350 mAh g−1 (0.1 A g−1; 1.0−3.5 V vs. Li/Li+) and a high capacity retention of about 81% after 3000 cycles (at a specific current of 2 A g−1 and 298 ± 2 K) in the organic electrolyte, demonstrating greatly enhanced structural stability of α-MoO3 within the heterostructure. Finally, we successfully fabricated an all-solid-state thin-film lithium microbattery based on the α-MoO3/η-Mo4O11 electrode, lithium phosphorous oxynitride (LIPON) solid electrolyte, and Li metal electrode. Benefiting from the stable positive electrode–solid electrolyte interface during Li+ intercalation/deintercalation processes, the microbattery showed a long-term cycling stability with about 74% capacity retention after 4000 cycles (at a specific current of 2 A g−1 and 298 ± 2 K).

Results

Construction of α-MoO3/η-Mo4O11 heterostructure and structural characterizations

The fabrication of α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films on Pt/Ti/glass and Pt/Ti/stainless steel substrates through direct current (DC) magnetron sputtering under O2/Ar atmosphere is schematically illustrated in Fig. 2a. During the sputtering process, the as-formed ionized Ar (Ar+) are driven towards the Mo target under the electric field. After Ar+ bombardment, the target Mo atoms are released to react with the O at ~400–800 K and deposit on the substrate in the form of Mo oxides19. X-ray diffraction (XRD) and corresponding Rietveld refinement were carried out to identify phase evolution and crystal structure of the films under different deposition conditions. The O2/Ar flow ratios of 8%, 13%, and 25% at a deposition temperature of 573 K were selected to prepare α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films, respectively, which are demonstrated by the powder XRD and glancing incidence XRD (GIXRD) patterns (Supplementary Figs. 1 and 2). Specifically, via Rietveld refinement, Fig. 2b reveals that the heterostructure possesses approximately 63 wt% α-MoO3 and 37 wt% η-Mo4O11, respectively, indicating α-MoO3 is the major phase for the heterostructure. The refined phase fraction error for η-Mo4O11 in the heterostructure is 1.3% as provided by the MAUD. The Rietveld-refined XRD patterns demonstrate a monoclinic structure for η-Mo4O11 (Supplementary Fig. 3, JCPDS No. 13–0142) and an orthorhombic structure for α-MoO3 phase (Fig. 2c and Supplementary Table 1, JCPDS No. 05–0508)13,20. Moreover, the chemical composition and Mo valence state of the film can be well controlled by tuning the oxygen partial pressure and substrate temperature during deposition. As shown in Fig. 2d, when the O2/Ar flow ratio was larger than 20%, Mo atoms can be completely oxidized to form α-MoO3. With the decrease of O2/Ar flow ratio to less than 20%, low valence molybdenum oxides such as η-Mo4O11 were gradually generated together with α-MoO3, resulting in the formation of α-MoO3/η-Mo4O11 heterojunction. Such heterojunction design can be easily achieved by controlling the O2/Ar flow ratio (10–20%) and substrate temperature (573–673 K). The standard deviation obtained during refinement for α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 is shown in Supplementary Table 2. The value in the bracket shows the standard deviation21. The phase evolution in thin films deposited under different O2/Ar flow ratios is further confirmed by Raman measurements (Supplementary Fig. 4). The Raman spectrum of the MoO3/η-Mo4O11 film was found to contain all the characteristic Raman bands of both α-MoO3 and η-Mo4O11, demonstrating the successful fabrication of the heterostructure in the film by magnetron sputtering22,23.

Fig. 2: Fabrication of the α-MoO3/η-Mo4O11 heterostructure.
figure 2

a Schematic diagram illustrating the fabrication of η-Mo4O11, α-MoO3/η-Mo4O11, and α-MoO3 thin films by magnetron sputtering. b, c Rietveld-refined XRD patterns of the α-MoO3/η-Mo4O11 and α-MoO3 thin films. The asterisk represents the substrate. d The corresponding O2/Ar flow ratio versus the deposition temperature phase diagram. eg Cross-section FESEM images of η-Mo4O11, α-MoO3/η-Mo4O11, and α-MoO3 thin films, respectively.

The surface morphologies of various samples were investigated by field-emission scanning electron microscopy (FESEM). As shown in Supplementary Fig. 5a–c, the increment of oxygen partial pressure not only induces the phase transition from η-Mo4O11 to α-MoO3 but also increases the surface roughness of the film. At a low O2/Ar flow ratio, the film exhibits a highly dense and smooth surface morphology, whereas at a high O2/Ar flow ratio, the film is composed of nanoflakes, resulting in a rough surface morphology. Therefore, increasing the O2/Ar flow ratio can enlarge particle size, change phase composition, and enhance the valance state of Mo ions of the films. The cross-section FESEM images of α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films are presented in Fig. 2e–g, which reveal that the corresponding thicknesses of different films are about 1.1, 1.0, and 1.3 µm, respectively.

To investigate the heterostructure and interface between α-MoO3 and η-Mo4O11, high-resolution transmission electron microscopy (HRTEM) and advanced spherical aberration-corrected scanning transmission electron microscopy (STEM) were employed for structural analysis. Ideal structure models of α-MoO3 and η-Mo4O11 superimposed on the high-angle annular dark field (HAADF)-STEM images were plotted to better illustrate the intergrowth of α-MoO3 and η-Mo4O11 phases and the formation of heterojunction. The atomic-resolution HAADF-STEM images of α-MoO3 (Fig. 3a) and η-Mo4O11 (Fig. 3b) samples reveal lattice spacing of ~0.69 nm and 0.28 nm, respectively, which correspond to (020) plane of layered α-MoO3 and (020) plane of η-Mo4O11. Noted that the distances between Mo atoms along [010] direction are not all the same due to the bilayer structure of α-MoO3. The corresponding lattice spacing of (020) plane of α-MoO3 thus can be classified as the intralayer and interlayer spacing. The structures shown in Fig. 3a, b agree well with Mo atomic arrangements of the orthorhombic MoO3 along [101] zone axis and monoclinic Mo4O11 along [100] zone axis (Fig. 3c). Supplementary Fig. 6 shows a low-magnification HAADF-STEM image of the α-MoO3/η-Mo4O11 sample, revealing that the thin film is composed of nanograins with irregular shapes. An interfacial region in the α-MoO3/η-Mo4O11 heterostructure is shown in Fig. 3d, in which η-Mo4O11 nano-domains with diameters of ~3–5 nm are found to be embedded in the layered α-MoO3 matrix (Fig. 3d and Supplementary Fig. 7). The atomic arrangement of Mo within the blue square agrees well with the crystal model of monoclinic η-Mo4O11 along the [100] zone axis, whereas the atomic arrangement of Mo within the red square conforms to the crystal model of the layered α-MoO3 along the [101] zone axis (Fig. 3d). Typically, the α-MoO3/η-Mo4O11 heterostructure demonstrates a specific orientation relationship of (020)α//(020)η and [10–1]α//[001]η with a coherent interface between α-MoO3 and η-Mo4O11 (Fig. 3d). Generally, the direct joining between α-MoO3 (020) and η-Mo4O11 (020) planes (or a sharp interface) will generate a large lattice mismatch (f) of ~18.8%, defined by f = (dα−2dη)/dα, which is energetically unfavorable to form a coherent interface24. However, the HAADF-STEM image at the interfacial region demonstrates that a gradual transition zone with a thickness of ~4–5 nm instead of a sharp interface separates the two phases, thereby enabling the gradual release of misfit strain. In contrast to the sharp interface, the gradual transition interface enables the gradual release of misfit strain between the two lattices, leading to a continuous lattice mismatch less than 5% at the interfacial region. Therefore, the gradual transition interface is critical to reducing lattice mismatch between α-MoO3 and η-Mo4O11, resulting in the construction of coherent interface between these two phases. Simultaneously, the large stress and strain issues associated with sharp interface can be effectively mitigated by the gradual transition interface design, further improving the structural stability of the heterostructure.

Fig. 3: Interface for the α-MoO3/η-Mo4O11 heterostructure.
figure 3

a HAADF-STEM image for layered α-MoO3 along [101] zone axis. The purple spheres represent Mo atoms and the yellow spheres represent oxygen atoms. b HAADF-STEM image for monoclinic η-Mo4O11 along [100] zone axis. c Crystal structures of orthorhombic MoO3 along [101] zone axis and monoclinic Mo4O11 along [100] zone axis. d HAADF-STEM image for the α-MoO3/η-Mo4O11 heterostructure. The blue square corresponds to the monoclinic Mo4O11 viewed along the [100] zone axis while the red square corresponds to the orthorhombic MoO3 viewed along the [101] zone axis. e The O K-edge ELNES spectra of the α-MoO3/η-Mo4O11 heterostructure obtained along the yellow arrow in the left STEM image.

Electron energy-loss spectroscopy (EELS) as a sensitive tool was employed to probe the oxidation state of Mo in the heterostructure. As shown in Fig. 3e, along the yellow line in the HAADF-STEM image, a series of energy-loss near-edge structures (ELNES) spectra of the O K-edge were recorded. The ELNES of O K-edge represents the transition of oxygen s core states to oxygen p states hybridized with Mo 4d orbitals25,26,27. In Mo–O octahedral configuration, the crystal field splitting leads to the generation of t2g and eg symmetry bands, corresponding to peak A at about 532 eV and feature B at about 536 eV, respectively. For α-MoO3, both the t2g and eg orbitals are empty, and thus O 1 s electrons can be excited into both orbitals. Consequently, the corresponding two peaks can be clearly detected at the O K-edge ELNES spectrum of MoO3. When the t2g orbitals in η-Mo4O11 are filled with electrons, an intensity decrease in peak A can be observed in O K-edge spectrum25. From the series of O K-edge ELNES spectra on the marked sites of α-MoO3/η-Mo4O11 in Fig. 3e, it can be observed that the spectra of 3 and 6 show obviously decreased intensity at peak A, suggesting embedded nano-domains of η-Mo4O11 in α-MoO3 matrix. The evolution of Mo chemical state in different samples was also investigated by X-ray photoelectron spectroscopy (XPS), as shown in Supplementary Fig. 8. Based on the Mo 3d XPS spectra, the α-MoO3 sample mainly contains Mo6+, and the trace of Mo5+ suggests the existence of oxygen vacancy. In comparison, the Mo5+/Mo6+ ratio notably increases in the α-MoO3/η-Mo4O11 sample as half Mo are 5+ in η-Mo4O11, further confirming the formation of heterostructure.

Electrochemical performance

To compare the electrochemical behaviors, the α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films were investigated in Li metal coin cells with excess Li and an electrolyte of lithium perchlorate (LiClO4) in a mixture of ethylene carbonate and dimethyl carbonate. All the cells were charged and discharged between 1.0 and 3.5 V (vs. Li/Li+) at a specific current of 0.1 A g−1. As shown in Fig. 4a, the initial charge/discharge curves of the Li//α-MoO3 half-cell show one pair of major potential plateaus at about 2.4 V (vs. Li/Li+) with a large reversible capacity of ~300 mAh g−1. Notably, the first discharge curve of the α-MoO3 electrode displays a minor potential plateau at ~2.8 V vs. Li/Li+, corresponding to the small anodic peak in cyclic voltammetry (CV) curve in Supplementary Fig. 9, which can be attributed to the irreversible phase transition to the new phase of LixMoO3 with Li+ entering the intralayer10,13,14,28. With such irreversible phase transition, the α-MoO3 electrode exhibits fast capacity decay even within five cycles. As shown in Fig. 4b, the charge/discharge curves of the Li//η-Mo4O11 half-cell present two pairs of potential plateaus between 1.5 and 3.5 V vs. Li/Li+ (corresponding to the two pairs of redox peaks in CV curves in Supplementary Fig. 9), delivering a large reversible capacity of about 350 mAh g−1. With α-MoO3 as the major phase in the heterostructure, the charge/discharge curves of the α-MoO3/η-Mo4O11 electrode display similar potential plateaus as the α-MoO3 electrode but with a more sloping profile (Fig. 4c). A high reversible capacity of ~350 mAh g−1 was achieved and well maintained by the α-MoO3/η-Mo4O11 electrode for the initial cycles. Surprisingly, the first discharge curve of the α-MoO3/η-Mo4O11 electrode maintains a steep slope at ~2.8 V vs. Li/Li+ without the appearance of potential plateau, suggesting successful suppression of the irreversible phase transition in the layered α-MoO3 within heterostructure during lithiation. To further identify the phase transition during the lithiation/delithiation of these Mo-based electrodes, we carried out differential capacity analysis (dQ/dV) by numerical differentiation of data in Fig. 4a–c29,30. As shown in Supplementary Fig. 10, the features at ~2.8 V vs. Li/Li+ in dQ/dV plot of α-MoO3 are absent in its 5th cycle, suggesting the irreversible phase transition of the α-MoO3 electrode. In sharp contrast, no irreversible phase transition is found in dQ/dV plot of α-MoO3/η-Mo4O11 electrode, which maintains the pristine redox peaks well in its 5th cycle.

Fig. 4: Electrochemical behavior of the α-MoO3/η-Mo4O11 heterostructure at 298 ± 2 K.
figure 4

ac Galvanostatic charge/discharge profiles of the α-MoO3, η-Mo4O11, and α-MoO3/η-Mo4O11 electrodes at the 1st, 2nd, and 5th cycles, respectively. d Rate capabilities of the α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 electrodes. e Cycle performance comparison of the α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 electrodes at 0.5 A g−1. f Long-term cycle performance of the α-MoO3/η-Mo4O11 electrode at 2.0 A g−1. Inset shows the charge–discharge curves at different cycle numbers.

Typical charge/discharge curves of the α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 electrodes at specific current from 0.1 A g−1 to 5 A g−1 are shown in Supplementary Fig. 11, and the corresponding rate capabilities of the three electrodes are compared in Fig. 4d. Layered α-MoO3, possessing a large interlayer distance, could favor fast Li+ diffusion through the two-dimensional diffusion channels. The rate performance of α-MoO3, however, is compromised by its irreversible phase transition and fast structural degradation, which slow down its electrode kinetics. When α-MoO3 is stabilized in the heterostructure, the α-MoO3/η-Mo4O11 electrode exhibits both large specific capacity and high rate performance. At a high specific current of 5 A g−1, the corresponding capacity retention of the α-MoO3/η-Mo4O11 electrode (with reference to the capacity at 0.1 A g−1) is ~41%, which is larger than ~31% of α-MoO3 and ~16% of η-Mo4O11.

The electrochemical impedance spectroscopy (EIS) spectra of the three electrodes were measured at open-circuit potentials and are shown in Supplementary Fig. 12. It can be observed that the charge transfer resistance of the α-MoO3/η-Mo4O11 electrode calculated by fitted EIS is the smallest among the three electrodes, as shown in Supplementary Fig. 12 and Supplementary Table 3, further demonstrating the improved charge transfer efficiency for the heterostructure. Meanwhile, the chemical diffusion coefficient of lithium (\({D}_{{{Li}}^{+}}\)) of α-MoO3, η-Mo4O11, and α-MoO3/η-Mo4O11 electrodes were calculated (Supplementary Fig. 13) by conducting the galvanostatic intermittent titration (GITT) technique during the lithiation process. The \({D}_{{{Li}}^{+}}\) values in α-MoO3 electrode are substantially higher than those in η-Mo4O11 electrode owing to the large ionic diffusion channels. However, the irreversible phase transition results in the dramatic decrease of the diffusion coefficient of Li+ in α-MoO3 electrode. Compared with both α-MoO3 and η-Mo4O11 electrodes, α-MoO3/η-Mo4O11 electrode possesses fast and stable Li+ diffusion, which could be assigned to the solid-state interfacial chemistry and improved structural stability.

Based on the ultraviolet photoelectron spectroscopy (UPS) and ultraviolet-visible absorption spectroscopy measurements (Supplementary Fig. 14a), the schematic energy-level diagrams of α-MoO3, η-Mo4O11, and heterostructured α-MoO3/η-Mo4O11 were proposed as shown in Supplementary Fig. 14b. The Fermi levels of both α-MoO3 and η-Mo4O11 are all located near their Conduction Bands, suggesting their n-type semiconductor feature. As α-MoO3 exhibits smaller work function (4.28 eV) as compared to that of η-Mo4O11 (4.58 eV), electrons tend to transfer from α-MoO3 to η-Mo4O11 across the interface. The electron flow leads to the accumulation of positive charge on the α-MoO3 side and negative charge on the η-Mo4O11 side near the interface. Simultaneously, the energy levels of α-MoO3 shift upward whereas those of η-Mo4O11 bend downward near the interface until their Fermi levels reach equilibrium31. Thus, a built-in electric field with a direction pointing from α-MoO3 to η-Mo4O11 is formed at the α-MoO3/η-Mo4O11 heterojunction. Given that η-Mo4O11 domains are randomly embedded within the α-MoO3 matrix, Li+ ions need to be transferred from α-MoO3 to η-Mo4O11 during the discharge process. During the discharge process, the as-formed built-in electric field can accelerate Li+ diffusion from α-MoO3 to η-Mo4O11, thus improving the electrode kinetics for the discharge process. Under the built-in electric field, Li+ could accumulate at η-Mo4O11 side to neutralize the negative charges, and the electric field around the hetero-interface may finally vanish after the charge balance32. Upon the charge process, Li+ ions are first extracted from α-MoO3 owing to its direct contact with electrolyte. As α-MoO3 possesses a more open structure for Li+ diffusion, the fast leaching of Li+ in α-MoO3 will lead to a Li+ concentration gradient across the α-MoO3/η-Mo4O11 interface, thus generating a reversed built-in electric field during the charge process. Under such an electric field, the transfer of Li+ from η-Mo4O11 to α-MoO3 can be accelerated, improving the electrode kinetics for charge process33. Overall, the constructed α-MoO3/η-Mo4O11 hetero-interface is beneficial to improve Li+ intercalation and deintercalation in this heterostructured electrode during both charge and discharge processes.

The structural stabilities of different electrodes were further investigated by cycling tests. Figure 4e compares the cycle performances of the three electrodes between 1.0 and 3.5 V (vs. Li/Li+) at a specific current of 0.5 A g−1. It is clear that the α-MoO3 electrode exhibits fast capacity fading with only approximately 27% capacity retained after 300 cycles, suggesting continuous structural degradation during cycling. In comparison, the η-Mo4O11 electrode presents superior structural stability with about 82% capacity retained after 300 cycles, demonstrating that it can be used as good structural stabilizer for the heterostructure design. Nevertheless, the specific capacity of the η-Mo4O11 electrode at 0.5 A g−1 is the lowest among the three electrodes due to its poor rate capability. Remarkably, the α-MoO3/η-Mo4O11 electrode displays both the highest specific capacity and the best cycle performance with a negligible capacity loss after 300 cycles. The superior structural stability of the heterostructure can be further demonstrated by a long-term cycling test. When cycled at a low specific current of 0.5 A g−1, the α-MoO3/η-Mo4O11 electrode can retain over 90% of the initial capacity after 1000 cycles. In addition, as depicted in Supplementary Fig. 15a, b, the redox peaks of α-MoO3/η-Mo4O11 electrode are well retained after 1000 cycles and the corresponding dQ/dV plots at the 1st and 1000th cycles are nearly overlapped with minimal capacity loss, further demonstrating the high stability of the heterostructure with coherent interface. In sharp contrast, the α-MoO3 displays an irreversible capacity loss of ~73% after 300 cycles at a specific current of 0.5 A g–1. Moreover, the redox peaks in dQ/dV plot of α-MoO3 are remarkably reduced and new peaks emerge even after 300 cycles, suggesting severe structural change.

As the α-MoO3/η-Mo4O11 electrode was subjected to prolonged cycling at 2 A g−1 (Fig. 4f), the initial reversible capacity is ~170 mAh g−1 and about 81% of the initial capacity can be retained after 3000 cycles, substantially outperforming previously reported α-MoO3-based positive electrodes (Supplementary Table 4), including H2O molecule modified α-MoO3, oxygen-deficient α-MoO3−x, α-MoO3−x/carbon nanotube, and pre-intercalated α-MoO3 (Supplementary Fig. 16)10,13,15,23,34,35. Importantly, the common phenomenon of fast capacity decay during initial cycles for the α-MoO3-based electrode as reported in literature is not observed in the present study for the α-MoO3/η-Mo4O11 heterostructured electrode, suggesting greatly enhanced structural stability of α-MoO3 within the heterostructure. Supplementary Fig. 17a–d presents the FESEM images of the α-MoO3 and α-MoO3/η-Mo4O11 thin films collected after different cycle numbers at 0.5 A g−1, respectively. It is noted that the morphology of the α-MoO3/η-Mo4O11 thin film was well maintained after 1000 cycles; however, severe morphology changes along with nanoparticles formation and aggregation, and cavities formation were observed in the α-MoO3 thin film after only 5 cycles (Supplementary Fig. 18). The photo images of different electrodes after 300 cycles in Supplementary Fig. 19 visually confirm the excellent stability of the α-MoO3/η-Mo4O11 thin film during the cycling process. In strong contrast, drastic exfoliation of the active material from the α-MoO3 thin film electrode was detected after 300 cycles, indicating large volume change and electrode pulverization occurred during the cycling.

Lattice pinning via coherent interface

To grasp insight into the mechanism of the excellent structural stability of the α-MoO3/η-Mo4O11 heterostructure, we have conducted ex situ XRD measurements on different electrodes at various charge and discharge states to elucidate the potential-dependent structure evolution and lattice variation (Fig. 5a, b). For α-MoO3, the gradual disappearance of the (020) peak belonging to the pristine α-MoO3 along with the emerging (030) peak of the lithiated α-MoO3 (LixMoO3, x~0.25), at ~2.8 V vs. Li/Li+, confirms the irreversible phase transformation during the initial discharge process10,13,14. Simultaneously, the intensive (110) and (040) peaks ascribed to α-MoO3 vanish and a new broad peak appears in the following lithiation process, indicating the original stacking order of Mo–O octahedron layers becomes strongly disordered28. In contrast to α-MoO3, all of the characteristic peaks for the α-MoO3/η-Mo4O11 heterostructure are well retained with only reversible peak shift during charge/discharge, demonstrating α-MoO3 layered structure can be stabilized without phase transition in the heterostructure. Meanwhile, the corresponding interlayered spacing of (020) planes of α-MoO3 in heterostructure increases from 6.80 Å at 3.5 V to 6.96 Å at 1.0 V (expansion rate ~2%), whereas the initial interlayer spacing of the pure α-MoO3 exhibits a considerably increased expansion rate of ~16% (from 6.80 to 7.90 Å), demonstrating the coherent interface can effectively pin lattice expansion of α-MoO3 in heterostructure. Unlike α-MoO3, the monoclinic η-Mo4O11 presents minimal and reversible lattice change during both lithiation and delithiation processes (Supplementary Fig. 20), suggesting that it could be the perfect structural stabilizer to constrain lattice variation of α-MoO3. It should be noted that the coherent interface not only suppresses the lattice variation of α-MoO3 but also further reduces the lattice variation of η-Mo4O11 in the heterostructure. As compared between Fig. 5b and Supplementary Fig. 20, the diffraction peak shift of η-Mo4O11 in heterostructure during charge and discharge is even smaller than that of the pure η-Mo4O11.

Fig. 5: Structure evolution upon Li+ intercalation/deintercalation.
figure 5

a, b Ex situ XRD measurements of the α-MoO3 and α-MoO3/η-Mo4O11 electrodes during the first discharge–charge processes in a voltage range between 1.0 and 3.5 V (vs. Li/Li+). η respresnts η-Mo4O11. c Mo 3d XPS spectra of α-MoO3 and α-MoO3/η-Mo4O11 at the delithiated state (3.5 V vs. Li/Li+). d, e Mechanism of structural degradation in α-MoO3 and enhanced structural stability of α-MoO3 induced by lattice pinning in α-MoO3/η-Mo4O11 during lithiation/delithiation processes. The purple spheres represent Mo atoms, the yellow spheres represent oxygen atoms, and red spheres represent lithium atoms.

The ex situ XPS measurements were performed to investigate the electrochemical reversibility of the α-MoO3/η-Mo4O11 heterostructure (Fig. 5c). When fully delithiated at 3.5 V, the α-MoO3/η-Mo4O11 heterostructure contains only Mo6+ and Mo5+, whereas the pure α-MoO3 presents ~18% Mo4+ in addition to Mo6+ and Mo5+, which suggests that more inactive Li+ are trapped in the pure α-MoO3 associated with the initial irreversible phase transition. Another interesting feature is that the fraction of Mo4+ at the full lithiation state in α-MoO3/η-Mo4O11 (~52%) is obviously higher than that in α-MoO3 (~41%; Supplementary Fig. 21), suggesting that the enhancement in specific capacity of the heterostructure originates from the enhanced redox activity of Mo6+ to Mo4+. The atomic structure of the cycled α-MoO3/η-Mo4O11 sample, which underwent 100 cycles at a specific current of 0.5 A g−1, was further analyzed by HADDF-STEM. As seen in Supplementary Fig. 22, both α-MoO3 and η-Mo4O11 phases connected by the coherent interface can be clearly observed. The Mo atomic arrangements in the red square agree well with the orthorhombic α-MoO3 viewed along [100] zone axis, while the Mo atomic arrangements in the yellow square agree well with η-Mo4O11 viewed along [001] zone axis, demonstrating that the coherent interface between α-MoO3 and η-Mo4O11 is highly stable during the cycling process. Schematic Fig. 5d illustrates that the origin of the irreversible phase transition and structure degradation in α-MoO3 lie in the large lattice expansion during lithiation process. In sharp contrast, as illustrated in Fig. 5e, the epitaxial lattice of α-MoO3 is well constrained by the coherent interface due to the strong ionic bonds between α-MoO3 and η-Mo4O11. With the strong lattice pinning effect, the layered structure of α-MoO3 is perfectly stabilized in the heterostructure, allowing efficient Li+ intercalation and deintercalation with minimal lattice variation.

Insights from theoretical calculations

DFT calculations were performed to gain in-depth insights into the improved structure stability of layered α-MoO3 induced by the lattice pinning effect. f-Li0.25MoO3 was used as a prototype for the lithiated state to compare with the pristine α-MoO3 because the irreversible phase transition occurs at this lithiation state of LixMoO3 (x = 0.25). A hypothetical c-Li0.25MoO3 structure with fixed interlayered spacing along the b axis as the pristine α-MoO3 was used to simulate the pinning effect on the suppression of lattice expansion. Supplementary Fig. 23a–c shows the relaxed lattice structures of α-MoO3, f-Li0.25MoO3, and c-Li0.25MoO3, respectively. An increased interlayer spacing of 7.9 Å along the b axis in f-Li0.25MoO3 was observed as compared to 6.8 Å in α-MoO3, which is in good agreement with the experimental result. Simultaneously, the distorted Mo6+ ions in MoO6 octahedra of α-MoO3 was found to move to the octahedral center after lithiation (Fig. 6a–c). It is well known that the off-center displacement of d0 Mo in MoO6 octahedra of α-MoO3 (Mo6+) was ascribed to the second-order Jahn–Teller (SOJT) effect through the interaction between the empty d-orbitals (d0) of Mo cations and filled p-orbitals of oxygen ions36,37,38,39. The formation of Mo5+/Mo4+ associated with Li+ intercalation into the layered α-MoO3, however, breaks this distortion state, resulting in the migration of Mo ions back to the center of MoO6 octahedra40. Therefore, lithiation in pristine α-MoO3 induces large lattice expansion together with Mo migration, which could promote the irreversible phase transition with poor structural stability (Supplementary Fig. 24).

Fig. 6: Insights from theoretical calculations.
figure 6

a, b Differential charge density of MoO3 and f-Li0.25MoO3. The blue and red regions represent the depletion and accumulation of charge, respectively. c Layered MoO3 with Mo off-centering and the displacement of Mo ion being shifted toward the orbital center during lithiation. d Differential charge density of c-Li0.25MoO3. e α-MoO3 with constrained lattice eliminates the migration of Mo ions and phase transition during lithiation.

Interestingly, the Mo ions in MoO6 octahedra of c-Li0.25MoO3 retain the same position as the pristine α-MoO3 (Fig. 6d), suggesting the migration of Mo ions is driven by lattice expansion during lithiation. By suppressing the interlayer expansion along the b direction, the migration barrier of Mo ions is greatly enhanced during the discharge process, which effectively inhibits the displacement of Mo ions to the center of MoO6 octahedra. In addition, partial density of states (PDOS) analysis indicates enhanced energy for the occupied Mo/O states in f-Li0.25MoO3 as compared to that in c-Li0.25MoO3, demonstrating weakened Mo–O interaction stemming from increased interlayered spacing in f-Li0.25MoO3 (Supplementary Fig. 25). As indicated in Supplementary Fig. 26, there are three types of oxygen atoms, i.e., Ot, Oa and Os, in the pristine α-MoO3. It can be seen that the Mo–Ot bonds in f-Li0.25MoO3 elongate from 2.372 to 2.461 Å, while the length of Mo–Ot bond in c-Li0.25MoO3 only changes from 2.372 to 2.371 Å (Supplementary Table 5) after lithiation. The notably increased Mo–Ot bonds greatly increase the probability for Mo migration in α-MoO3, thus enabling the irreversible phase transition during the initial lithiation. Based on the DFT calculations, it is speculated that lattice pinning via the coherent interface can significantly increase the structural stability of α-MoO3 in the heterostructure, successfully inhibiting the adverse phase transition (Fig. 6e).

All-solid-state thin-film microbattery

Owing to the solid−solid electrode /electrolyte interface in all-solid-state lithium batteries, the positive electrode materials with large volume change during charge/discharge are detrimental to interface stability, which could result in fast capacity fading and even battery failure41. Herein, the heterostructured α-MoO3/η-Mo4O11 electrode with minimal lattice variation during cycling is anticipated to find desirable application in solid-state lithium batteries42,43. To identify the compatibility of the α-MoO3/η-Mo4O11 electrode with solid electrolyte, we fabricated an all-solid-state thin-film lithium microbattery using α-MoO3/η-Mo4O11 as the positive electrode, LiPON as the solid electrolyte, and metal lithium as the negative electrode (Fig. 7a and Supplementary Fig. 27). The suitable ionic conductivity (3 × 10−6 S cm−1; 298 ± 2 K) and low electric conductivity (1 × 10−10 S  cm−1; 298 ± 2 K) of LiPON with high stability make it an appropriate solid electrolyte for solid-state thin-film lithium batteries19,35,44. The LiPON was sputtered using a Li3PO4 target under N2 atmosphere at 298 K. During the RF magnetron sputtering process, the introduced N2 gas was split into nitrogen ions and electrons under the high electric field, resulting in a reactive nitrogen cation species. The ionized N species can react with the Li-P-O species to form LiPON45,46. As shown in the cross-section FESEM image of the solid-state battery (Fig. 7b), the LiPON affords close contact with both the α-MoO3/η-Mo4O11 electrode and lithium metal electrode, guaranteeing efficient charge transfer at the interface with low interfacial impedance. In addition, the STEM image and corresponding energy-dispersive X-ray spectroscopy (EDS) elemental mappings are shown in Supplementary Fig. 28 and Fig. 7c, respectively. Well-separated and uniform distributions of Mo and P can be clearly observed in the sample, further demonstrating the intimate interface contact.

Fig. 7: All-solid-state thin-film lithium microbattery.
figure 7

a Schematic diagram illustrating the configuration of all-solid-state thin-film lithium microbattery. b The cross-section FESEM image of the thin-film microbattery. c The cross-section STEM image and corresponding EDS elemental mappings of the thin-film microbattery. d Rate performance of the thin-film microbattery. e Cycle performance of the flexible thin-film microbattery at different bending states at 0.05 A g−1. f Long-term cycle performance of thin-film microbattery at 2 A g−1. The solid electrochemical energy storage tests were carried out at 298 ± 2 K.

The charge/discharge curves of the solid-state thin-film lithium microbattery based on the α-MoO3/η-Mo4O11 positive electrode at different specific currents are shown in Supplementary Fig. 29. At a low specific current of 0.05 A g−1, the microbattery delivers a large capacity of ~300 mAh g−1 (~0.075 mAh cm−2) at the first cycle (Fig. 7d). Even at a high specific current of 2 A g−1, it can still exhibit ~90 mAh g−1, suggesting good rate performance. The specific capacity and rate performance of the α-MoO3/η-Mo4O11 electrode in the solid-state thin-film battery are relatively lower than those of the α-MoO3/η-Mo4O11 electrode tested in organic electrolyte, suggesting limited charge transport kinetics in solid-state battery configuration. Nevertheless, the as-fabricated solid-state microbattery device is highly flexible and robust (Fig. 7e). The cycling test of the device showed negligible capacity change under various bending angles from 0 to 180°. Even after cycling at different bending states, the interface between α-MoO3/η-Mo4O11 electrode and LiPON electrolyte maintains the intimate contact, suggesting the good flexibility of the solid-state microbattery and potential application as power sources for various flexible electronics (Supplementary Figs. 30 and 31). Importantly, Fig. 7f shows the long-term cycling performance of the solid-state microbattery, presenting ~74% capacity retention even after 4000 cycles at 2 A g−1. The long-term cycling performance of the α-MoO3/η-Mo4O11 electrode within a solid-state microbattery is well-placed among previously reported solid-state microbatteries (Supplementary Fig. 32)19,35,45, demonstrating the substantially enhanced battery cycle life in solid-state configuration when the α-MoO3/η-Mo4O11 heterostructure is employed.

Based on the Galvanostatic charge/discharge profile in Supplementary Fig. 29, the α-MoO3/η-Mo4O11-based thin film micorbattery delivers a specific energy density of ~0.48 mWh with an areal energy density of 0.16 mWh cm–2. Moreover, the α-MoO3/η-Mo4O11 based solid-state microbattery can achieve superior electrochemical performance in comparison with the LiCoO2 and LiMn2O4-based solid-state microbatteries, especially in specific capacity, max energy density, and cycle life19,44. Importantly, for the α-MoO3/η-Mo4O11 and α-MoO3 thin film deposition, a cheap Mo metal target can be directly used in sputtering, making the fabrication cost much lower than LiCoO2 and LiMn2O4 thin films. In addition, in contrast to traditional Li-contained positive electrode thin films (LiCoO2 and LiMn2O4) that necessitate a high annealing temperature around 973 K, the α-MoO3/η-Mo4O11 as well as α-MoO3 thin films can be well crystallized at a much lower annealing temperature of 673 K, making battery on-chip integration compatible without damaging the chip.

Discussion

In summary, we demonstrate a lattice pinning strategy via constructing coherent interface to stabilize α-MoO3 in the intergrown α-MoO3/η-Mo4O11 heterostructure. This strategy boosts Li+ storage with substantially prolonged cycle life and improved rate capability. With the rational heterostructure design, the epitaxial lattice of α-MoO3 is effectively constrained by η-Mo4O11 at the coherent interface via the strong ionic bonds, which successfully suppresses lattice expansion along b direction and eliminates the adverse phase transition of α-MoO3 during the lithiation process. The DFT calculations indicate that lithiation-induced interlayer expansion in α-MoO3 weakens the Mo–O bonds and allows Mo migration in MoO6 octahedra, which could drive the irreversible phase transition to form the unstable LixMoO3 phase. With minimal lattice variation during lithiation/delithiation, the α-MoO3/η-Mo4O11 heterostructure affords substantially extended cycle life without the fast capacity loss within initial cycles as compared to individual α-MoO3. The high structural stability associated with minimal lattice variation of the α-MoO3/η-Mo4O11 heterostructure is further demonstrated in all-solid-state thin-film lithium microbattery, which presents a long-term cycling stability with about 74% capacity retention after 4000 cycles at 2 A g−1. This study demonstrates effective lattice pinning in layered α-MoO3 via coherent interface design, and provides new insights to design highly stabilized intercalation hosts for sustainable batteries.

Methods

Materials synthesis

The Mo target is industrially available material, which was purchased from Wuxi Kaistar Electro-optic Material Co., LTD. The α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films were deposited on Pt/Ti/glass or Pt/Ti/stainless steel substrates using a pure Mo target (75 mm in diameter and 5 mm in thickness) through DC magnetron sputtering (SKY Technology Development Co., Ltd, China) under O2 and Ar atmosphere. Before deposition, a background pressure less than 1 × 10−5 Pa was first reached in the chamber. The distance between target and substrate was kept at 10 cm and the substrate rotation is 20 rpm. The DC power was fixed at 60 W. The α-MoO3 thin film was deposited at 573 K, in a working pressure of ~1 Pa introduced by mass flow controllers with 40.0 sccm Ar and 10.0 sccm O2 (25% O2/Ar ratio). α-MoO3/η-Mo4O11 thin film was deposited at 573 K, in a working pressure of ~1 Pa introduced by mass flow controllers with 40.0 sccm Ar and 5.0 sccm O2 (13% O2/Ar ratio). η-Mo4O11 thin film was deposited at 573 K, in a working pressure of ~1 Pa introduced by mass flow controllers with 40.0 sccm Ar and 3.0 sccm O2 (8% O2/Ar ratio). All the films were deposited for 1.5 h and then annealed in the chamber at 673 K for 1 h to obtain the final α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films. The average thickness of α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films are about 1.3, 1.0, and 1.1 µm, respectively. The mass loadings of the α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films determined by a Sartorius Analytical Balance (CPA225D, with resolution of 10 µg) were 0.21, 0.25, and 0.30 mg cm−2, respectively.

Materials characterizations

The crystallographic information of the samples was characterized by XRD (Bruker) with Cu Kα radiation, Raman spectroscopy (Jobin-Yvon T6400 Micro-Raman system), XPS (Phi Quantera SXM spectrometer equipped with an Al Kα X-ray excitation source). For the Rietveld refinement process with the materials analysis using diffraction (MAUD) software, data is first refined according to the background and instrument parameters. Then, we used the structural model of α-MoO3 (a = 3.95 Å, b = 13.83 Å, c = 3.69 Å, Z = 4, space group Pbnm, ICSD no. 36167) as the starting model in the refinement. Moreover, “excess” peaks that cannot be indexed to α-MoO3 (ICSD no. 36167) were further refined using the structural model of η-Mo4O11 (a = 24.54 Å, b = 5.44 Å, c = 6.70 Å, Z = 4, space group P21/a, ICSD no. 15814), generating a low-reliability factor Rwp. Specifically, the quantitative value (64/36) of weight fraction between α-MoO3 and η-Mo4O11 is extracted from the MAUD based on the following equation (Eq. (1))47,48:

$${W}_{i}=\frac{{S}_{i}{({ZMV})}_{i}}{\mathop{\sum }\limits_{j=1}^{N}{S}_{j}{({ZMV})}_{j}}$$
(1)

where V, M, Z, and S are the unit-cell volume of phase i, the molecular weight, the number of molecules per unit cell, and the scale factor in the α-MoO3/η-Mo4O11 sample, respectively. Considering the thin film texture effect, Rietveld refinement on thin film XRD patterns was carried out by using the harmonic texture model implemented into the MAUD software49,50,51. For the electronic structure, we used UPS (Thermo ESCALAB 250XI), which provides valence band and work function position. The morphologies and microstructures of the samples were investigated by FESEM (FEI Quanta 250 F) and HRTEM (FEI Tecnai G2). A JEM-ARM200F STEM fitted with a double-aberration corrector for both probe-forming and imaging lenses was utilized to perform the HAADF/ABF imaging and EELS; the electron microscope was operated at 200 kV. The attainable resolution of the probe defined by the objective pre-field was 78 pm.

Electrochemical measurements

Li coin cells using the α-MoO3, α-MoO3/η-Mo4O11, and η-Mo4O11 thin films as the positive electrodes, 1 M LiClO4 in ethylene carbonate and dimethyl carbonate (EC: DMC, 1:1, v/v) solution as the electrolyte (~100 μL), Celgard 2400 as the separator (16 mm in diameter and 25 μm in thickness), and lithium foils (99.9% in purity) as negative electrodes were assembled into R2025-type coin cells in an argon-filled glove box (O2 < 0.1 ppm, H2O < 0.1 ppm). The porosity and average pore size of the separator are 41%, and 0.043 μm, respectively. Moreover, the lithium foil has a diameter of 14.5 mm and a thickness of 500 μm.

The \({D}_{{{Li}}^{+}}\) is calculated by GITT according to the following equation (Eq. (2)):

$${D}_{{GITT}}=\frac{4}{\pi \tau }{\left(\frac{{n}_{m}{V}_{m}}{S}\right)}^{2}{\left(\frac{{\Delta E}_{s}}{{\Delta E}_{t}}\right)}^{2}$$
(2)

where nm represents the mole value, τ is relaxation time, and S represents the area of the electrode, Vm is the mole volume, ∆Et represents the changed voltage during the discharge, and ∆Es is the voltage response under the pulse current.

The all-solid-state thin-film battery was fabricated through a layer-by-layer deposition procedure. Stainless steel foil with a Ti/Pt transition layer was used as the substrate, and the α-MoO3/η-Mo4O11 thin film positive electrode was first deposited by DC sputtering. The LiPON solid electrolyte film and lithium metal film were then sequentially deposited onto the α-MoO3/η-Mo4O11 thin film (Supplementary Fig. 27). Typically, the LiPON electrolyte film with a thickness of 2.5 μm was deposited onto the positive electrode by RF reactive magnetron sputtering (SKY Technology Development Co., Ltd, China) using a Li3PO4 target (75 mm in diameter and 5 mm in thickness) under N2 atmosphere at 298 K for 15 h. Then, the lithium metal layer of approximately 2.5 μm was deposited on the LiPON layer by thermal evaporation (SKY Technology Development Co., Ltd, China). It should be noted that the area of LiPON is larger than those of the positive electrode and negative electrode to avoid a short circuit at the edge. After the deposition of 500 nm Cu current collector by RF reactive magnetron sputtering (SKY Technology Development Co., Ltd, China) using a Cu target, the all-solid-state thin-film microbattery was finally sealed with aluminum-plastic film. Galvanostatic charge–discharge measurements were operated in the voltage range between 1.0 and 3.5 V (vs. Li/Li+) at different specific currents using a Neware BTS4000 battery test system. CV measurements were conducted on an electrochemical workstation (Biologic VSP). EIS measurements were carried out by applying a potentiostatic amplitude of 10 mV over the frequency range of 100.0 kHz to 0.1 Hz on the Biologic VSP. Six data points are collected per decade of frequency. The open-circuit voltage time applied before carrying out the EIS measurement is ~10 h. Zview software is used to fit the EIS and calculate the errors between the raw and fitted data. The electrochemical energy storage tests were carried out at 298 ± 2 K.

DFT calculation

First-principles calculations based on DFT were performed to understand the effect of Li-ion intercalation on the crystalline structure of α-MoO3 using the Vienna Ab initio Simulation Package (VASP). The projector-augmented wave (PAW) method was used to describe the ion-electron interactions. Perdew–Burke–Ernzerhof (PBE) version of generalized gradient approximation (GGA) was adopted for the exchange-correlation energy. A kinetic energy cutoff of 520 eV was used for the plane wave expansion of the valence electron wave functions. A dense Γ-centered Monkhorst-Pack k-point mesh with a sampling density of 0.04 Å−1, 10−6 eV/cell in total energy, and 10−2 eV/Å in force were adopted for the convergence criterion during structural optimization. Because of the layer structure of MoO3, van der Waals density functional (vdWDF) of optB86b-vdW functional was performed during structural optimization. Differential charge density was calculated by the difference between the total valence charge density of the structure and the superposition of the valence charge densities of neutral atoms.